Está en la página 1de 14

Abdullah H. AlEssa and Mohammed Q.

Al-Odat

ENHANCEMENT OF NATURAL CONVECTION HEAT


TRANSFER FROM A FIN BY TRIANGULAR
PERFORATIONS OF BASES PARALLEL AND TOWARD
ITS BASE
Abdullah H. AlEssa and Mohammed Q. Al-Odat*
Al-Balqa Applied University, Al-Huson University College
Al-Huson, Irbid, Jordan

1. INTRODUCTION
The removal of excessive heat from system components is essential to avoid the damaging effects of burning or
overheating. Therefore, the enhancement of heat transfer is an important subject of thermal engineering. The heat
transfer from surfaces may in general be enhanced by increasing the heat transfer coefficient between a surface and
its surroundings, by increasing the heat transfer area of the surface, or by both. In most cases, the area of heat
transfer is increased by utilizing extended surfaces in the form of fins attached to walls and surfaces [1]. Extended
surfaces (fins) are frequently used in heat exchanging devices for the purpose of increasing the heat transfer between
a primary surface and the surrounding fluid. Various types of heat exchanger fins, ranging from relatively simple
shapes, such as rectangular, square, cylindrical, annular, tapered, or pin fins, to a combination of different
geometries, have been used [2]. One of the primary goals in the design of modern thermal systems is the
achievement of more compact and efficient heat exchangers. Reducing energy loss due to ineffective use and
enhancement of energy transfer in the form of heat has become an increasingly important duty for the design
engineers of thermal systems, considering the world wide increase in energy demand. This duty requires employing
heat transfer surfaces with high heat transfer coefficients and high area compactness. Particular attention in the
selection/design of heat transfer surfaces is required if the energy carrying fluid selected is a gas [3]. Fins as heat
transfer enhancement devices have been quite common. As the extended surface technology continues to grow, new
design ideas emerge, including fins made of anisotropic composites, porous media, and perforated and interrupted
plates [1–6]. Due to the high demand for lightweight, compact, and economical fins, the optimization of fin size is of
great importance. Therefore, fins must be designed to achieve maximum heat removal with minimum material
expenditure, taking into account, however, the ease of manufacturing of the fin shape [4–8]. Large number of studies
has been conducted on optimizing fin shapes. Other studies have introduced shape modifications by cutting some
material from fins to make cavities, holes, slots, grooves, or channels through the fin body to increase the heat
transfer area and/or the heat transfer coefficient [7–9]. One popular heat transfer augmentation technique involves
the use of rough or interrupted surfaces of different configurations. The surface roughness or interruption aims at
promoting surface turbulence that is intended mainly to increase the heat transfer coefficient rather than the surface
area [1,4,5,6,9]. It was reported that non-flat surfaces have free convection coefficients that are 50% to 100% more
than those of flat surfaces [1,9]. Several other researchers reported a similar trend for interrupted, perforated, and
serrated surfaces, attributing the improvement to the restarting of the thermal boundary layer after each interruption,
indicating that the increase in convection coefficient is even more than enough to offset lost area, if any [1].

* Corresponding author:
P.O. Box 50, Al-Huson, Jordan
Al-Huson University College
Al-Balqa Applied University
Irbid, Jordan
E-mail: m_alodat@yhaoo.com

Paper Received February 27, 2008; Paper Revised April 4, 2009; Paper Accepted May 27, 2009

October 2009 The Arabian Journal for Science and Engineering, Volume 34, Number 2B 531
Abdullah H. AlEssa and Mohammed Q. Al-Odat

Perforated plates (fins) represent an example of surface interruption [1,7–10] and are widely used in different
heat exchangers, film cooling, and solar collector applications. This study is aimed mainly at examining the extent of
heat transfer enhancement from a horizontal rectangular fin under natural convection conditions as a result of
introducing body modification (perforations) to the fin body. The modification in this work is vertical equilateral
triangular perforations made through the fin thickness. The study investigates the influence of perforations on heat
transfer rate or heat dissipation rate of the perforated fin. The modified fin (perforated fin) is compared to the
corresponding solid (non-perforated) fin in terms of heat transfer rate. The study eventually attempts to make the
best use of the material and size of a given fin, which involves some sort of optimization. The specific objectives of
the work may be summarized as follows:
1- Evaluate the influence of relevant fin and perforation factors on the enhancement of heat transfer rate.
2- Determine the values of parameters that would result in maximum heat transfer enhancement of the perforated
fin compared to the solid counterpart.
2. HEAT TRANSFER ANALYSIS
The classical treatment of fins, which assumes one-dimensional heat conduction as the Biot number is small (less
than 0.01), can be considered [11,12]. The perforated fin with triangular perforations analyzed in this study is shown
in Figure 1. Figure 2 shows the symmetry part considered for heat transfer analysis (shown hatched). For this part of
the fin, the transverse Biot number in (z) direction ( Bi z ) can be expressed ( Bi z = h pc.t / 2k ) and the transverse Biot
number in (y) direction ( Bi y ) can be calculated by ( Bi y = h .(S + b/2)/k ). As the values of ( Bi z ) and ( Bi y ) are
ps y
less than 0.01, then the heat transfer in (z) and (y) directions can be assumed lumped and a one dimension solution
can be considered. If the values of ( Bi z ) and ( Bi y ) are greater than 0.01, then the heat transfer solution must be two
or three dimensions. In this study, the parameters of the perforated fin are taken as they lead to values of ( Bi z ) and
( Bi y ) smaller than 0.01. The analysis and results reported are based on the following assumptions:

1- Steady, one-dimensional heat conduction


2- Homogeneous and isotropic fin material with constant thermal conductivity
3- No heat sources/sinks in the fin body
4- Uniform base and ambient temperatures
5- Side area of the fin is much smaller than that of its surface area (w >>t)
6- Uniform heat transfer coefficient overall the fin surfaces (perforated or solid)
7- Negligible radiation effects.

z y

b t
2Sx

2Sy

L W

Figure 1. The fin with equilateral triangular perforations (perforation bases are parallel and toward the fin base)

532 The Arabian Journal for Science and Engineering, Volume 34, Number 2B October 2009
Abdullah H. AlEssa and Mohammed Q. Al-Odat

Sy

2Sy

Sx 2Sx Sx
x
x

Figure 2. The symmetrical hatched part used in the mathematical formulation of the perforated fin
Based on the above assumptions, the energy equation of the fin along with the boundary conditions may be stated
as below [13]:
d 2T
k =0 (1)
dx 2
The associated boundary conditions are
1- At the fin base (x = 0):-
T =T (2)
b
2- At the perforated surfaces.
dT
k L + h (T − T ) + h (T − T ) = 0 (3)
dx x ps ∞ pc ∞

As shown in Figure 1, the heat transfer surface area including the tip of the fin with triangular perforations is
expressed as
A = ( 2W .L − 2N .N .A ) + (W . t) + (N .N .A )
pf x y c x y pc
= A + N .N (A − 2A )
f x y pc c
= A + N .N .b( 3t − bSin( 60o )/ 2 ) (4)
f x y

The material weight of the perforated fin is compared with that of the non-perforated one by weight ratio RWF,
which is given by
RW F =W /W
pf sf
= (L.W . t − N .N .A .t) / (L.W .t)
x y c
= 1- (N .N .A .t) / (L.W .t)
x y c
= 1 - N .N .b 2Sin( 60o ) / ( 2L.W ) (5)
x y

The numbers of triangular perforations ( N x and N y ) in the perforated fin for a given length and width are
obtained from the following expressions:

October 2009 The Arabian Journal for Science and Engineering, Volume 34, Number 2B 533
Abdullah H. AlEssa and Mohammed Q. Al-Odat

N = Int(L/( 2S
+ b. Sin( 60o ))) (6)
x x
N = Int(W /( 2S + b) (7)
y y

In the expressions above the term “Int” is an integer number function. In this study, the energy Equation (1) is
solved numerically utilizing the finite-element technique. The variational approach in matrix notation [13] is used in
formulating the algebraic equations of the problem of this perforated fin. The corresponding variational statement
has the following form [13]:
1 dT 2 1
I = ∫∫∫ k ( ) dV e + ∫∫ h ps (T − T ∞ ) dA ps
2
n 2V e dx 2A
ps
1
+ ∫∫ h pc (T − T ∞ ) dA pc + ∫∫ ht (T t − T ∞ )T dA t
2
(8)
2A A
pc t

The rationale for solving the heat transfer problem of a fin with triangular perforations may be best clarified by
reference to Figures 1, 2, and 3. Figure 2 shows the symmetry part (shown hatched) that is considered for analysis.
Figure 3 considers in more details the part under study and shows the existence of three regions (A, B, and C) around
any perforation In this case, the regions A and C are uniform, whereas the region B is tapered. The regions A and C
are divided into ( N f ) elements each, while the region B is divided into ( N t ) elements each. The values of ( N f ) and
( N t ) are arbitrary according to the mesh generation requirements. The total number of finite elements ( N e ) and the
total number of their nodes ( N n ) are expressed as

N = N ( 2N + 2N ) (9)
e x f t
N = N +1 (10)
n e

y
A B

1 2

3 4 5 A
C
6 7 B
x

Figure 3. Expanded symmetrical part with the three regions A, B, and C, considered in the mathematical formulation for the fin
with triangular perforation. (1, 2, 3, ...... N e ) are numbers of the linear finite elements
The results of the solution of Equation (8) are the perforated fin temperature distribution along its length
direction (x coordinate). Once the temperature distribution along the perforated fin length is obtained, the heat
dissipation rate from the perforated fin ( Q ) can be computed by the following expression [14]:
pf
N ⎛T +T
Q = 2N . ∑ ⎜ I
e I + 1 −T ⎞⎟ ⎛ h ⎛ Pe(I) + Pe(I + 1) ⎞ Le(I) + h .L (I).t ⎞ + Q + Q (11)
pf y ⎜ 2 ∞ ⎟ ⎜⎝ ps ⎜⎝ 2

⎠ pc pc ⎟ t
⎠ s
I = 1⎝ ⎠
It is established that everything else being the same, heat dissipation from a fin, solid or perforated, depends on
fin surface area and heat transfer coefficient. For the solid fin, both aspects are established. The average value of
( hss ) is that for a single horizontal plate in natural convection and may be given by [15]
h = Nu .k / Lc (11)
ss air
Lc = L.W /( 2L + 2W ) (12)

534 The Arabian Journal for Science and Engineering, Volume 34, Number 2B October 2009
Abdullah H. AlEssa and Mohammed Q. Al-Odat

The average Nusselt number, Nu, is given by [15]

Nu + Nu
Nu = u l (13)
2
Nu = [( 1.4 / ln( 1 + 1.4 / ( 0.43Ra 0.25 )))10 + ( 0.14Ra 0.333 )10 ] 0.1 (14)
u
Nu = 0.527Ra 0.2/( 1 + ( 1.9/ Pr ) 0.9 ) 2/ 9 (15)
l
6 8
where (10 < Ra < 10 ).

For the fin considered in this study, both perforated and non-perforated, the fin tip is a vertical surface for which
the Nusselt number ( Nu t ) is given by [15]
( 1/ 6 )
⎛⎛ ⎞
6

⎜⎜ ⎟ ⎟
⎜⎜
( )
6⎟
2 .8 ⎟ + .103Ra 0.333
Nu = 0.5 ⎜ ⎜ ⎟ ⎟ (16)
t ⎛ 2.8 ⎞
⎜ ⎜ ln 1 + ⎟ ⎟
⎜⎜ ⎜ ⎟
.25 ⎠ ⎟ ⎟
⎝ ⎝ ⎝ 0. 515 Ra ⎠ ⎠
h = Nut . k / Lc where Lc = L. t/( 2L + 2t) (17)
t air
As for the perforated fin, the aspect of surface area was discussed in the previous section. However, three distinct
heat transfer coefficients exist. The first one is the heat transfer coefficient of the remaining solid portion of the
perforated surface ( h ps ) which can be calculated by the following expression [4]:

h = ( 1 + 0.75 ROA).h (18)


ps ss
N .N .b 2 .Sin( 60o )
OA x y
ROA = = (19)
OA L.W
max
The second one is the heat transfer coefficient within the perforation ( h pc ). Its Nusselt number ( Nu c ) is given

by [15]
1/ - 1.5
⎡⎛ Ra ⎞-1.5 ⎤

Nu = ⎜ c ⎟ + ( 0.62Ra 0 . 25 ) -1. 5 ⎥ (20)
c ⎢⎜ 13.3 ⎟ c ⎥
⎣⎢ ⎝ ⎠ ⎦⎥
Ra = g.β. (T −T ).L 4 /(t.ν.µ) (21)
c m ∞ c
L = 2A /P (22)
c c c
h = Nu .k /L (23)
pc c air c
where Ra c < 104
The third one is the heat transfer coefficient at fin tip ( ht ), the Nusselt number ( Nu t ) of which is given by
Equations (16 and 17). The ratio of heat dissipation of the perforated fin to that of the solid one (RQF) is introduced
and given by

R QF = Q / Q (24)
pf sf

The heat dissipation rate of the perforated fin ( Q pf ) is computed according to the finite element heat transfer
solution described in [10]. The temperature distribution along the non-perforated (solid) fin, including tip
temperature (Tt) and heat dissipation rate ( Q ), are, respectively, expressed and computed according to the exact
sf
solution described in Reference [16], as shown in the following equations:

October 2009 The Arabian Journal for Science and Engineering, Volume 34, Number 2B 535
Abdullah H. AlEssa and Mohammed Q. Al-Odat

T -T Cosh(m (L − x)) + (h /(m ∗ k)) Sinh(m (L - x))


x ∞ = t (25)
T −T Cosh(m ∗ L) + (h /(m ∗ k)) Sinh(m ∗ L))
b ∞ t
T −T
T =T + b ∞ (26)
t ∞ Cosh(m ∗ L) + (h /(m ∗ k)) Sinh(m ∗ L)
t
Sinh(m ∗ L) + (h /(m ∗ k)) Cosh(m ∗ L)
Q = k ∗ A ∗ m(T −T ) t (27)
sf b ∞ Cosh(m ∗ L) + (h /(m ∗ k)) Sinh(m ∗ L)
t
W here (m) is defined as
h ∗P
m= f (28)
k ∗A

3. RESULTS AND DISCUSSION


It is believed that comparing the perforated fin with its solid counterpart is the best means to evaluate the
improvement or otherwise in heat transfer brought about by introducing the perforations. It is assumed that both fins
have the same dimensions (the fin length is L = 50 mm and its width is W = 100 mm), same thermal conductivity,
and same base and ambient temperatures, T b and T ∞ of 100 and 20 oC, respectively. In this section, the findings
relate to temperature distribution along the fin and heat transfer rate. The effect of introducing the perforations on
these aspects is examined by comparing the perforated fin to its solid counterpart.
3.1. Temperature Distributions
The temperature distributions along the perforated fin, Tpf(x), is investigated in terms of perforation parameters
and fin thickness where the perforation spacing is (Sx=Sy=1 mm). The temperature distribution along the fin length is
plotted in Figure 4. As indicated in Figure 4, it is obvious that the temperature distributions show non-uniform
curves caused by perforations. The perforations create a variation in sectional area along the fin length and then lead
to a variation in the fin thermal resistance. The effect of variation of sectional area of the perforated fin on its thermal
resistance decreases as the thermal conductivity increases, so the curves become more uniform. Figure 4 shows that
the temperature difference (temperature drop) between the fin base and its tip increases as the triangular perforation
dimension (b) is increased. This is because the thermal resistance of the perforated fin increases as (b) is increased.
Consequently, from a temperature distribution viewpoint, it is recommended to use as small as possible perforation
dimensions. Also, it can be deduced from temperature distribution that fin temperatures increase as the fin thickness
is increased. This is readily explained by the fact that the thermal resistance of the perforated fin decreases as the fin
thickness is increased. Therefore, from the temperature distribution viewpoint, it is preferable to use fin thicknesses
as large as possible.

536 The Arabian Journal for Science and Engineering, Volume 34, Number 2B October 2009
Abdullah H. AlEssa and Mohammed Q. Al-Odat

100

99

95
Tpf [°C]

94
90

t=4 [mm]
85 t=2 [mm]
b=7 [mm]
b=7 [mm] 89

0 10 20 30 40 50 0 10 20 30 40 50

x [mm] x [mm]

k=500 [W/m.°C]
k=400 [W/m.°C]
k=300 [W/m.°C]
k=200 [W/m.°C]
k=100 [W/m.°C]
Figure 4. Temperature distribution along the perforated fin length with variable fin thickness and perforation dimension

3.2. Ratio of Heat Dissipation Rate (RQF)


The ratio of heat dissipation rate from the perforated fin to that of the corresponding solid one (RQF) is studied in
terms of perforation dimension (b) for different values of fin thickness (t) and thermal conductivity (k). Figure 5
shows the variations of (RQF) with the perforation dimension (b) for different values of thermal conductivity (k)
ranging from 40 to 500 [W/m.oC] and 3 mm thickness. The variation of (RQF) with (b) at various (k) showed a
regular trend of increasing to a maximum limit then it declines. This trend can be explained by taking into
consideration the net effects of changing in fin heat transfer surface area and heat transfer coefficients due to
perforations. Furthermore, the figure shows that (RQF) has an obvious relation with the perforation dimension (b).
The perforation dimension at which the (RQF) ratio has its maximum value is referred to as the optimum perforation
dimension ( bo ). Also, it can be noted that the fin thermal conductivity has a slight influence on the optimum
perforation dimension ( bo ). Figure 6 presents the variation of ( bo ) under a wide range of (k) (i.e., 50 to 500 W/m
o
C) for different values of fin thickness from 1 to 5 mm. It is clear that the thickness has a dominant effect on (bo) in
comparison with that of (k). This indicates that the geometrical influence is much more than that of the thermo-
physical properties of the fin material.

October 2009 The Arabian Journal for Science and Engineering, Volume 34, Number 2B 537
Abdullah H. AlEssa and Mohammed Q. Al-Odat

1.8 t=3[mm]
Sx=1[mm]
Sy=1[mm]
1.6
RQF

1.4
k=500 [W/m.°C]
k=300 [W/m.°C]
1.2 k=100 [W/m.°C]
k=60 [W/m.°C]
k=40 [W/m.°C]
1.0
1 3 5 7 9 11

b [mm]

Figure 5. The ratio of heat dissipation rate distribution with respect to perforation dimension at various fin thermal conductivities

7
bo[mm]

5 t=5[mm]
Sx=1[mm] t=4[mm]
t=3[mm]
Sy=1[mm] t=2[mm]
t=1[mm]
3

0 100 200 300 400 500

k [W/m.°C]
Figure 6. The optimum perforation dimension ( bo ) distribution with respect to the fin thermal conductivity at various fin
thicknesses

538 The Arabian Journal for Science and Engineering, Volume 34, Number 2B October 2009
Abdullah H. AlEssa and Mohammed Q. Al-Odat

4.2. The Effect of Longitudinal Spacing

To investigate the effect of the perforation longitudinal spacing ( S x ) on the perforated fin performance, the heat
dissipation ratio (RQF) is plotted as a function of the spacing ( S x ), as shown in Figure 7. This figure indicates that
at any t the (RQF) decreases with increasing of spacing ( S x ). This behavior is due to the fact that increasing of ( S x )
means smaller number of perforations, this means loss of the element that causes heat transfer enhancement. So
according to the ( S x ) spacing it is preferable to minimize it as possible.

k=500[W/m.°C], bo=7.9[mm]
k=300[W/m.°C], bo=7.8[mm]
k=100[W/m.°C], bo=7.7[mm]
1.8 k=60[W/m.°C], bo=7.6[mm]
k=40[W/m.°C], bo=7.5[mm]
RQF

1.6

t=3[mm]
1.4 Sy=1[mm]

1 2 3 4 5 6

Sx[mm]
Figure 7. The ratio of heat dissipation rate (RQF) distribution with respect to the longitudinal spacing ( S x ) at various fin
thermal conductivities

4.3. The Effect of Lateral Spacing

The effect of lateral spacing ( S y ) on the perforated fin performance is elucidated in Figure 8, such that the
variation of (RQF) with ( S y ) is studied for various values of the fin thickness and its thermal conductivity. It is
clear that RQF has a sudden sharp increase under low values of ( S y ) then tends to decline thereafter. However, low
values of ( S y ) means more perforation area and less solid material, which increase the fin thermal resistance, which
causes a reduction in RQF. Moreover, it can be said that the conflicting effects of fin thermal resistance and number
of perforations are responsible for this style of fin thermal behavior.

Figure 8 shows that (RQF) severely depends on the spacing ( S y ). It is obvious that there is an optimum value of
the spacing ( S yo ) at which RQF has its maximum value. The values of ( S yo ) strongly depend on the fin thickness
and its thermal conductivity, as shown in Figure 9. This figure shows that ( S yo ) decreases as fin thermal
conductivity and its thickness increase.

October 2009 The Arabian Journal for Science and Engineering, Volume 34, Number 2B 539
Abdullah H. AlEssa and Mohammed Q. Al-Odat

2.0 k=500 [W/m.°C]


k=300 [W/m.°C]
k=100 [W/m.°C]
k=60 [W/m.°C]
k=40 [W/m.°C]
1.8
RQF

1.6

t=3[mm]
Sx=1[mm]
1.4
b=bo[mm]

0 1 2 3 4

Sy[mm]

Figure 8. The ratio of heat dissipation rate (RQF) distribution with respect to the lateral spacing ( S y ) at various fin thermal

conductivities

Sx=1[mm] t=5[mm]
2.0 t=4[mm]
b=bo[mm]
t=3[mm]
t=2[mm]
1.5 t=1[mm]
Syo[mm]

1.0

0.5

0
100 200 300 400 500

k[W/m.°C]

Figure 9. The optimum lateral spacing ( S yo ) distribution with respect to the thermal conductivity of the perforated fin at various

fin thicknesses

540 The Arabian Journal for Science and Engineering, Volume 34, Number 2B October 2009
Abdullah H. AlEssa and Mohammed Q. Al-Odat

To summarize the advantages of using the perforated fin at the optimum values of perforation geometry, the
ratios of heat dissipation (RQF) and of weight ratio (RWF) are plotted as a function of the fin thickness with variable
fin thermal conductivity at the optimum perforation diameter ( bo ) and optimum lateral spacing ( S yo ), as shown in
Figures 10 and 11. Figures 10 and 11 show that the use of perforations in fins leads to enhanced heat dissipation and
at the same time decreases the fin weight. The enhancement of heat dissipation and the weight reduction increase as
the fin thickness and its thermal conductivity increase. This means that the use of perforated fin leads to enhanced
heat dissipation rates and decreases the expenditure of fin material.

2.0
RQF

1.5

t=5[mm]
t=4[mm]
1.0 Sx=1[mm] t=3[mm]
b=bo t=2[mm]
t=1[mm]
Sy=Syo
0.5
100 200 300 400 500

k[W/m.°C]
Figure 10. The ratio of heat dissipation rate (RQF) distribution with respect to the fin thermal conductivity at various fin thicknesses for optimum
perforation dimension ( bo ) and optimum lateral spacing ( S yo )

Sx=1[mm]
t=5[mm]
b=bo t=4[mm]
0.8
Sy=Syo t=3[mm]
t=2[mm]
t=1[mm]
RWF

0.7

0.6
100 200 300 400 500

k[W/m.°C]
Figure 11. The ratio of perforated fin weight (RWF) distribution with respect to the fin thermal conductivity at various fin
thicknesses for optimum perforation dimension ( bo ) and optimum lateral spacing ( S yo )

October 2009 The Arabian Journal for Science and Engineering, Volume 34, Number 2B 541
Abdullah H. AlEssa and Mohammed Q. Al-Odat

5. CONCLUSIONS
Natural convection heat transfer enhancement from a horizontal rectangular fin embedded with equilateral
triangular perforations is numerically investigated. The heat dissipation rate from the perforated fin is compared to
that of the equivalent solid one. The effect of geometrical dimensions of the perforated fin (L, W, b, Sx, and Sy) and
thermal properties of the fin is studied in detail. The major findings of this study can be summarized by the
following points:
1. For certain values of triangular dimensions, the perforated fin can result in heat transfer enhancement. The
magnitude of enhancement is proportional to the fin thickness and its thermal conductivity.
2. The extent of heat dissipation rate enhancement for perforated fins is a complicated function of the fin
dimensions, the perforation geometry, and the fin thermophysical properties.
3. The gain in heat dissipation rate for the perforated fin is a strong relation of both, the perforation dimension,
and lateral spacing. This relation attains a maximum value at given perforation diameter and lateral spacing,
which is called the optimum perforation dimension ( bo ) and optimum spacing ( S yo ), respectively.

4. The perforation of fins enhances heat dissipation rates and at the same time decreases the expenditure of the
fin material.
NOMENCLATURE
A: cross sectional area of the fin
A : cross sectional area of the perforation
c

B : Biot number, h.L c /k .


i
b: triangular perforation dimension
g: gravity acceleration, g = 9.81m/sec2.
h: heat transfer coefficient
k: thermal conductivity of fin material
L: fin length
l: unit vector
Lc : characteristic length

N: number of perforations
N e : total number of finite elements of the perforated fin

Nf : number of finite elements in one of the uniform regions (A or C)

N : total number of nodes.


n
N : number of finite elements in the tapered region (B)
t

Nu: average Nusselt number, h.L / k


c
Nu : average Nusselt number of the inner perforation surface
c
OA: open area of the perforated surface
Q: heat transfer rate

Ra: Rayleigh number, g.β. (T − T ).L 4 /(t.ν.µ)


m ∞ c
Ra : Rayleigh number of the perforation inner lining surface
c

ROA: ratio of open area, N .N .b 2.Sin (60o ) / L .W


x y

542 The Arabian Journal for Science and Engineering, Volume 34, Number 2B October 2009
Abdullah H. AlEssa and Mohammed Q. Al-Odat

RQF: Ratio of heat dissipation rate of the perforated fin to that of the non-perforated one, Q /Q
pf sf

RWF: ratio of the perforated fin weight to that of the solid fin (perforated fin weight ratio), W /W .
pf sf

S: perforation spacing
T: temperature
t: fin thickness
W: fin width
W : perforated fin weight
pf

W : solid fin weight


sf
Subscripts and superscripts
b: fin base
l: lower surface of the fin
m: mean value
max: maximum
pc: perforation inner surface (within the perforation)
pf: perforated fin
ps: perforated surface or the remaining solid portion of the perforated fin
sf: solid (non-perforated) fin
ss: solid surface
t: fin tip
u: upper surface of fin
x: longitudinal direction or coordinate
y: transverse (lateral) direction with the fin width or coordinate
z: transverse (lateral) direction with the fin thickness or coordinate
∞ : ambient
REFERENCES
[1] A.E. Bergles, “Technique to Augment Heat Transfer”, in Handbook of Heat Transfer Applications, Second Edition,
Ch. 3, eds. Werren M. Rohsenow, James P. Hartnett, and Ejup N. Ganic, NY: McGraw-Hill.
[2] Bayram Sahin and Alparslan Demir, “Performance Analysis of a Heat Exchanger Having Perforated Square Fins”,
Applied Thermal Engineering, 6(2008), pp. 621–632.
[3] Bayram Sahin and Alparslan Demir, “Thermal Performance Analysis and Optimum Design Parameters of Heat
Exchanger Having Perforated Pin Fins”, Energy Conversion and Management, available online 4 January 2008,
doi:10.1016/j.enconman. 2007.
[4] A. H. Al-Essa and F. M. S. Al-Hussien, “The Effect of Orientation of Square Perforations on the Heat Transfer
Enhancement from a Fin Subjected to Natural Convection”, Heat and Mass Transfer, 40(2004), pp. 509–515.
[5] R. Mullisen and R. Loehrke, “A Study of Flow Mechanisms Responsible for Heat Transfer Enhancement in
Interrupted-Plate Heat Exchangers”, Journal of Heat Transfer (Transactions of the ASME), 108(1986), pp. 377–385.
[6] C. F. Kutscher, “Heat Exchange Effectiveness and Pressure Drop for Air Flow Through Perforated Plates With and
Without Crosswind”, Journal of Heat Transfer, 116(1994), pp.391–399.
[7] B. V. S. S. S. Prasad and A. V. S. S. K. S. Gupta, “Note on the Performance of an Optimal Straight Rectangular Fin
With a Semicircular Cut at the Tip”, Heat Transfer Engineering, 14(1998).
[8] B. T. F. Chung and J. R. Iyer, “Optimum Design of Longitudinal Rectangular Fins and Cylindrical Spines with
Variable Heat Transfer Coefficient”, Heat Transfer Engineering, 14(1993), pp. 31–42.

October 2009 The Arabian Journal for Science and Engineering, Volume 34, Number 2B 543
Abdullah H. AlEssa and Mohammed Q. Al-Odat

[9] S. Kakac, A. E. Bergles, and F. Mayinger, Heat Exchangers, Thermal-Hydraulic Fundamentals and Design.
Hemisphere Publishing Corporation, 1981.
[10] A. H. Al-Essa, “Enhancement of Thermal Performance of Fins Subjected to Natural Convection Through Body
Perforation”, Ph.D. thesis, Department of Mechanical Engineering, University of Baghdad, Iraq, and University of
Science and Technology, Jordan, 2000.
[11] A. Aziz and V. Lunadini, “Multidimensional Steady Conduction in Convicting, Radiating, and Convicting
Radiating Fins and Fin Assemblies”, Heat Transfer Engineering, 16(1995), pp. 32–64.
[12] P. Razelos and E. Georgiou, “Two–Dimensional Effects and Design Criteria for Convective Extended Surfaces”,
Heat Transfer Engineering 13(3)(1992), pp. 38–48.
[13] S. S. Rao, “The Finite Element Method in Engineering”, Second Edition, Elmsford, NY: Pergamon, 1989.
[14] Abdullah H. M. AlEssa, “One-Dimensional Finite Element Heat Transfer Solution of a Fin with
Triangular Perforations of Bases Parallel and Towered Its Base”, Archive of Applied Mechanics,
Accepted: 17 June 2008, DOI 10.1007/s00419-008-0250-5).
[15] G. D. Raithby and K. G. T. Holands, “Natural Convection”, in Handbook of Heat Transfer Applications, Second
Edition, Ch. 6, eds. W. Rohsenow, J. Hartnett, and E. Ganic , New York: McGraw-Hill, 1984.
[16] Frank P. Incropera and David P. Dewitt, “Fundamentals of Heat and Mass Transfer”, Fourth Edition, New
York: John Wiley and Sons, 1996, p. 110.

544 The Arabian Journal for Science and Engineering, Volume 34, Number 2B October 2009

También podría gustarte